19
An automated system for continuous measurements of trace gas fluxes through snow: an evaluation of the gas diffusion method at a subalpine forest site, Niwot Ridge, Colorado Brian Seok Detlev Helmig Mark W. Williams Daniel Liptzin Kurt Chowanski Jacques Hueber Received: 18 September 2008 / Accepted: 24 February 2009 / Published online: 6 May 2009 Ó The Author(s) 2009. This article is published with open access at Springerlink.com Abstract An experimental system for sampling trace gas fluxes through seasonal snowpack was deployed at a subalpine site near treeline at Niwot Ridge, Colorado. The sampling manifold was in place throughout the entire snow-covered season for continuous air sampling with minimal disturbance to the snowpack. A series of gases (carbon dioxide, water vapor, nitrous oxide, nitric oxide, ozone, volatile organic compounds) was determined in interstitial air withdrawn at eight heights in and above the snowpack at *hourly intervals. In this paper, carbon dioxide data from 2007 were used for evaluation of this technique. Ancillary data recorded inlcuded snow physical properties, i.e., temperature, pressure, and density. Various vertical concentration gradients were determined from the multiple height measurements, which allowed calculation of vertical gas fluxes through the snowpack using Fick’s 1st law of diffusion. Comparison of flux results obtained from different height inlet combinations show that under most conditions fluxes derived from individual gradient intervals agree with the overall median of all data within a factor of 1.5. Winds were found to significantly influence gas concentration and gradi- ents in the snowpack. Under the highest observed wind conditions, concentration gradients and calcu- lated fluxes dropped to as low as 13% of non-wind conditions. Measured differential pressure amplitude exhibited a linear relationship with wind speed. This suggests that wind speed is a sound proxy for assessing advection transport in the snow. Neglecting the wind-pumping effect resulted in considerable underestimation of gas fluxes. An analysis of depen- dency of fluxes on wind speeds during a 3-week period in mid-winter determined that over this period actual gas fluxes were most likely 57% higher than Electronic supplementary material The online version of this article (doi:10.1007/s10533-009-9302-3) contains supplementary material, which is available to authorized users. B. Seok D. Helmig M. W. Williams D. Liptzin K. Chowanski J. Hueber Institute of Arctic and Alpine Research (INSTAAR), University of Colorado at Boulder, Boulder, CO 80309-0450, USA B. Seok (&) Department of Atmospheric and Oceanic Sciences, University of Colorado at Boulder, Boulder, CO 80309-0311, USA e-mail: [email protected] M. W. Williams Department of Geography, University of Colorado at Boulder, Boulder, CO 80309-0260, USA D. Liptzin Department of Ecology and Evolutionary Biology, University of Colorado at Boulder, Boulder, CO 80309-0334, USA Present Address: D. Liptzin Department of Environmental Science, Policy, and Management, University of California at Berkeley, Berkeley, CA 94720-3114, USA 123 Biogeochemistry (2009) 95:95–113 DOI 10.1007/s10533-009-9302-3

An automated system for continuous measurements of trace ...snobear.colorado.edu/Markw/Research/09_seok.pdfcontinuous air sampling with minimal disturbance to the snowpack. A series

  • Upload
    others

  • View
    5

  • Download
    0

Embed Size (px)

Citation preview

  • An automated system for continuous measurements of tracegas fluxes through snow: an evaluation of the gas diffusionmethod at a subalpine forest site, Niwot Ridge, Colorado

    Brian Seok Æ Detlev Helmig ÆMark W. Williams Æ Daniel Liptzin ÆKurt Chowanski Æ Jacques Hueber

    Received: 18 September 2008 / Accepted: 24 February 2009 / Published online: 6 May 2009

    � The Author(s) 2009. This article is published with open access at Springerlink.com

    Abstract An experimental system for sampling

    trace gas fluxes through seasonal snowpack was

    deployed at a subalpine site near treeline at Niwot

    Ridge, Colorado. The sampling manifold was in place

    throughout the entire snow-covered season for

    continuous air sampling with minimal disturbance

    to the snowpack. A series of gases (carbon dioxide,

    water vapor, nitrous oxide, nitric oxide, ozone,

    volatile organic compounds) was determined in

    interstitial air withdrawn at eight heights in and

    above the snowpack at *hourly intervals. In thispaper, carbon dioxide data from 2007 were used for

    evaluation of this technique. Ancillary data recorded

    inlcuded snow physical properties, i.e., temperature,

    pressure, and density. Various vertical concentration

    gradients were determined from the multiple height

    measurements, which allowed calculation of vertical

    gas fluxes through the snowpack using Fick’s 1st law

    of diffusion. Comparison of flux results obtained

    from different height inlet combinations show that

    under most conditions fluxes derived from individual

    gradient intervals agree with the overall median of all

    data within a factor of 1.5. Winds were found to

    significantly influence gas concentration and gradi-

    ents in the snowpack. Under the highest observed

    wind conditions, concentration gradients and calcu-

    lated fluxes dropped to as low as 13% of non-wind

    conditions. Measured differential pressure amplitude

    exhibited a linear relationship with wind speed. This

    suggests that wind speed is a sound proxy for

    assessing advection transport in the snow. Neglecting

    the wind-pumping effect resulted in considerable

    underestimation of gas fluxes. An analysis of depen-

    dency of fluxes on wind speeds during a 3-week

    period in mid-winter determined that over this period

    actual gas fluxes were most likely 57% higher than

    Electronic supplementary material The online version ofthis article (doi:10.1007/s10533-009-9302-3) containssupplementary material, which is available to authorized users.

    B. Seok � D. Helmig � M. W. Williams �D. Liptzin � K. Chowanski � J. HueberInstitute of Arctic and Alpine Research (INSTAAR),

    University of Colorado at Boulder, Boulder,

    CO 80309-0450, USA

    B. Seok (&)Department of Atmospheric and Oceanic Sciences,

    University of Colorado at Boulder, Boulder,

    CO 80309-0311, USA

    e-mail: [email protected]

    M. W. Williams

    Department of Geography, University of Colorado at

    Boulder, Boulder, CO 80309-0260, USA

    D. Liptzin

    Department of Ecology and Evolutionary Biology,

    University of Colorado at Boulder, Boulder,

    CO 80309-0334, USA

    Present Address:D. Liptzin

    Department of Environmental Science, Policy, and

    Management, University of California at Berkeley,

    Berkeley, CA 94720-3114, USA

    123

    Biogeochemistry (2009) 95:95–113

    DOI 10.1007/s10533-009-9302-3

    http://dx.doi.org/10.1007/s10533-009-9302-3

  • fluxes calculated by the diffusion method, which

    omits the wind pumping dependency.

    Keywords Diffusion model � CO2 flux �Gradient method � Snowpack � Wind-pumping

    Introduction

    Recent research in snow-covered temperate, alpine,

    and arctic environments shows that subniveal (under

    the snow) soil respiration can result in substantial

    fluxes of carbon dioxide (CO2) through the snowpack

    (Sommerfeld et al. 1993; Brooks et al. 1996; Zimov

    et al. 1996; Oechel et al. 1997; Mariko et al. 2000;

    McDowell et al. 2000; Welker et al. 2000; Roehm and

    Roulet 2003; Monson et al. 2005, 2006). The winter-

    time CO2 flux through the snow can contribute in

    excess of 20% to the annual carbon budget (Winston

    et al. 1997). To date, studies of trace gas flux through

    snow have focused primarily on CO2. These findings

    suggest that microbial processes in underlying soil

    similarly may drive the flux of other gases such as nitric

    oxide (NO), nitrous oxide (N2O), and methane (CH4).

    Therefore, these processes warrant consideration in the

    annual ecosystem carbon (C) and nitrogen (N) budgets

    of seasonally snow-covered ecosystems (Sommerfeld

    et al. 1993; Brooks et al. 1996; VanBochove et al.

    1996; Mast et al. 1998; Yashiro et al. 2006).

    The ecological controls on the trace gas fluxes

    through the seasonal snowpack are still poorly

    known, in part because of the difficulty in measuring

    these fluxes and their dependencies during the winter

    season. The trace gases CO2, CH4, and N2O are

    greenhouse gases that play important roles in climate

    forcing. NO is a precursor of tropospheric ozone

    (O3), which is also a greenhouse gas, formed by

    photochemical reactions. Several studies have

    pointed out the high sensitivity of respiration, deni-

    trification, and other processes, which produce these

    gases, to soil temperature under the seasonal snow-

    pack. In turn, the soil temperature tends to increase

    with increasing snow depth because the insulating

    quality of snow cover increases with depth (Brooks

    and Williams 1999). Consequently, the depth and

    duration of the seasonal snowpack have been shown

    to have notable impacts on trace gas exchanges and

    ecosystem C and N budgets (Williams et al. 1998;

    Monson et al. 2006). For these reasons, there is a

    need to better quantify these processes, their behav-

    ior, and dependency on conditions such as snowpack

    properties, snow chemical composition, and soil

    biogeochemistry, in order to develop conceptual

    models for the description of these trace gas fluxes

    under current and future climate conditions.

    We have improved experimental protocols for the

    measurement of fluxes of trace gases through a

    seasonal snowpack over five winter seasons (2004–

    2008) at a high elevation subalpine site in the

    Colorado Rocky Mountains. Trace gases studied

    included CO2, nitrogen oxides (NOx = NO ? NO2),

    N2O, O3, and volatile organic compounds (VOC).

    The high resolution, multiple season data obtained

    were used to test the gas diffusion model (DM)

    method for deriving quantitative C and N winter

    balances. The DM approach has been the most

    common method for evaluating trace gas flux through

    snow (McDowell et al. 2000), as it has a number of

    advantages over other flux techniques such as cham-

    ber methods and eddy covariance measurements. In

    particular, the concentrations and gradients of trace

    gases in the seasonal snowpack are often substantially

    higher than ambient levels. This allows for more

    robust measurements and experiments appropriate for

    complex mountainous terrain and wintertime condi-

    tions. Furthermore, these snowpack gradient mea-

    surements can be conducted without meeting the

    stringent requirements for homogenous and flat

    terrain essential for eddy covariance measurements.

    The DM relies on the assumption that gas transfer

    inside the snowpack is determined by molecular

    diffusion. However, there are other factors such as

    pressure-related phenomena that influence gas

    exchange from the soil through any permeable

    medium (in particular, snow) to the atmosphere

    (Massman 2006). A number of studies (Massman

    et al. 1997; Jones et al. 1999; Hubbard et al. 2005;

    Takagi et al. 2005; Suzuki et al. 2006) pointed out the

    large potential error that may occur when the DM

    approach is applied to trace gas flux through the

    snowpack because of the effects of advection or

    wind-pumping. However, quantitative descriptions

    and corrections of this effect on the flux calculation

    remain highly uncertain.

    In this paper, we present the experimental

    approach that we have developed over the last

    5 years for all winter measurements of trace gas

    fluxes through a seasonal snowpack. Here, we focus

    96 Biogeochemistry (2009) 95:95–113

    123

  • on the flux of CO2 during 2007, but believe that this

    evaluation is applicable to other trace gases and other

    years. The decision to use CO2 for evaluating our

    trace gas sampling system was because the CO2record has been the longest and most complete of all

    gases studied to date, the relative precision and

    accuracy of its analytical measurement was highest,

    thus minimizing errors from the analytical determi-

    nation, and the CO2 data offered the highest number

    of comparison with studies reported in the literature.

    We evaluated potential errors in the estimation of

    CO2 fluxes through snow including sampling fre-

    quency, number of measurement heights within the

    snowpack, and physical properties of the snowpack

    such as the density of snow. Errors associated with

    the DM technique will be discussed with particular

    emphasis on the influence of wind-pumping. Results

    and discussion of measured gas fluxes and its

    ecological significance can be found in papers by

    Filippa et al. (2009, for N2O), Helmig et al. (2009, for

    NOx), and Liptzin et al. (2009, for CO2).

    Diffusion model (DM) measurement technique

    The DM method appeared to be most suitable for this

    study, because it allowed continuous long-term mea-

    surements throughout the winter season with relatively

    simple measurement techniques. This method pro-

    vided a direct approach to measuring the gas transport

    through the snowpack, since trace gases were sampled

    from inside the snow. In past research, gases from the

    snowpack were typically drawn from two levels,

    the soil–snow interface and the atmosphere above the

    snow surface, to obtain a concentration gradient (e.g.

    Jones et al. 1999; Takagi et al. 2005; Monson et al.

    2006; Suzuki et al. 2006). Besides the measurement of

    trace gases at different heights, the DM method

    depends on the physical properties of the porous

    medium. For the snowpack, this involves measure-

    ments of snow depth, density, and temperature. A

    newer technique not evaluated here involves measure-

    ments of gas concentration in the snowpack using

    nondispersive infrared (NDIR) sensors (Takagi et al.

    2005).

    The DM method is based on Fick’s 1st law of

    diffusion. The underlying theory assumes that the

    transport of a gas, e.g., CO2, in the snowpack operates

    primarily in the vertical dimension with fluxes dictated

    by the concentration differential. Within the horizontal

    footprint of the CO2 measurement, snow properties are

    assumed constant in each arbitrary layer of the

    snowpack, leading to consistent CO2 concentrations

    at each layer in the snowpack. Thus, CO2 efflux

    through the snowpack is calculated using the equation

    FCO2 ¼ �DCO2oCCO2

    oz

    � �; ð1Þ

    where FCO2 is the molecular flux of CO2(lmol m-2 s-1), DCO2 is the diffusivity of CO2 inthe snowpack airspace at the qz interval (mol m2 s-1),and oCCO2=oz is the CO2 concentration gradient in the

    snowpack (lmol m-3). Therefore, the flux of gas isdirectly proportional to the concentration gradient.

    This indicates that any changes in the concentration

    gradient term will directly change the flux. The

    diffusivity DCO2 can be estimated using the relation

    DCO2 ¼ /sDP0P

    T

    T0

    � �a; ð2Þ

    where / is the snowpack porosity, s is the tortuositycoefficient, D is the diffusion coefficient of the specific

    gas under consideration at standard temperature and

    pressure (T0 = 273.15 K and P0 = 1,013 hPa), i.e.,

    D = 0.1381 9 10-4 m2 s-1 for CO2. P is the ambient

    pressure (hPa), T is the snowpack temperature (K), and

    the exponent a = 1.81 is a theoretically set coefficientexplained by Massman (1998). Snowpack porosity /is calculated by

    / ¼ 1� qsnow=qiceð Þ; ð3Þ

    where qsnow is the measured density of snow at the qzinterval (kg m-3), and qice is the density of ice(917 kg m-3, Eisenberg and Kauzmann 1969). For

    the tortuosity constant s, Duplessis and Masliyah(1991) developed a theoretical relationship where the

    tortuosity can be derived from the porosity as

    s ¼ 1� 1� /ð Þ2=3

    /: ð4Þ

    However, we used

    s ¼ /1=3; ð5Þ

    which is a relationship that was empirically devel-

    oped from a snowpack in similar regions as our study

    site (Mast et al. 1998; Hubbard et al. 2005). Some

    researchers dealt with tortuosity by either ignoring it

    or by letting it be equal to 1 (e.g. Sommerfeld et al.

    Biogeochemistry (2009) 95:95–113 97

    123

  • 1993), or compensating for its effects by reducing the

    calculated flux value by 35% (e.g. Fahnestock et al.

    1998).

    Materials and methods

    Study site

    This study was conducted at the Soddie site on Niwot

    Ridge in an open meadow surrounded by ribbon forest

    just below tree line near the Continental Divide

    in the Colorado Rocky Mountains (40�0205200N;105�3401500W; 3,345 m above sea level). The site ison a 10� southwest facing slope (Erickson 2004). It hasan underground laboratory 3 9 9 9 2.4 m in size, line

    power, and an array of snow and zero-tension soil

    lysimeters (Williams et al. 2009). The soil in the open

    meadow is classified as a mixed Typic Humicryept

    (Soil Survey Staff 2006) and the forest vegetation is

    primarily a mixture of spruce (Picea englemannii) and

    fir (Abies lasiocarpa) [Williams et al. 2009]. The depth

    of A horizon is *34 cm; soil composition is 46%sand, 33% silt, 21% clay, and 31% organic matter in

    the upper 10 cm. The soil pH ranges between 4.7 and

    5.0 (Seastedt 2001).

    Trace gas sampling system

    Snowpack gas flux studies began in the 2003–2004

    winter season and have continued with additional gas

    and physical snowpack measurements. The introduc-

    tion paper in this issue by Williams et al. (2009)

    provides the history of the studies. In this manuscript,

    we report the system configuration from the 2006–

    2007 winter season (from here on referred to as

    winter 2007), because this season has the most

    extensive measurements to date. The photographs in

    Fig. 1 and the schematic in Fig. 2 depict the multi-

    inlet snow tower used in winter 2007. The snow

    tower was constructed of square aluminum alloy

    (3.8 9 3.8 cm) tubing with 60 cm long cross bars

    (1.3 9 1.3 cm) at heights of 0, 10, 30, 60, 90, 120,

    150, and 245 cm above the ground. The tower was

    installed in fall 2006, and snow was allowed to

    accumulate around it. The snowpack surrounding the

    tower was not disturbed throughout the winter season.

    Each of the eight cross bars supported a pair

    of sampling inlets. The inlets were fitted with

    25 mm Acrodisc� hydrophobic polytetrafluoroethyl-

    ene (PTFE) syringe filters (Pall Life Sciences, Ann

    Arbor, Michigan, USA) to prevent debris from

    entering the sampling line. Air was drawn at a rate

    of 1.5–3 L min-1 from each sampling height depend-

    ing on the number of chemical measurements con-

    ducted. Since the sampling flow was split between the

    paired inlets, the effective sampling rate per inlet was

    0.75–1.5 L min-1.

    The selection of a pair of inlets for sampling at a

    particular height on the tower was done through an

    array of eight solenoid valves. Sampling through each

    pair of inlets was done sequentially from the 245 cm

    height to the 0 cm height. Each sampling interval was

    10 min long, with trace gas contents determined in

    this gas flow every 10 s. From that record, 1-min

    averages were calculated for each trace gas, sampling

    interval, and measurement height. During the transi-

    tion from one sampling height to the next, concen-

    tration readings for both gases adjusted over *1 minto the new level at the lower inlet. A complete cycle

    took 80 min, and there were 18 cycles per day. The

    manifold used 2-way solenoid valves with PTFE

    body seals (Cole-Parmer, Vernon Hills, Illinois,

    USA). All the sampling lines and valves were

    conditioned over night with a flow of 2–3 L min-1

    of air containing 200–300 ppb ozone prior to instal-

    lation to minimize the loss of O3 in the manifold

    during subsequent field sampling. All sampling tubes

    were directed to the underground laboratory, which

    housed the analytical instruments. Sampling lines

    were all perfluoroalkoxy (PFA) Teflon�, inner diam-

    eter of 3.9 mm and outer diameter of 6.4 mm (Parker

    Hannifin, Cleveland, Ohio, USA), with equal lengths

    of 18 m. Sections of the sampling lines outside the

    laboratory were wrapped in pipe insulation with a

    self-controlling water pipe heater to maintain line

    temperatures slightly above 0�C. This was done toprevent water from freezing and clogging the sam-

    pling lines.

    Monitored gases included CO2, H2O, NOx, N2O,

    and O3. CO2 and H2O were measured using a LI-7000

    IRGA (LI-COR Environmental, Lincoln, Nebraska,

    USA). NOx and O3 were measured using a model

    42CTL chemiluminescence NOx analyzer, and model

    49 UV photometric O3 analyzer, respectively (Thermo

    Environmental Instruments, Franklin, Massachusetts,

    USA). N2O was measured by isothermal gas chroma-

    tography with electron capture detection (GC-ECD)

    98 Biogeochemistry (2009) 95:95–113

    123

  • [Shimadzu GC-8AIE, Shimadzu Scientific Instru-

    ments, Columbia, Maryland, USA]. On site instrument

    calibrations for the LI-7000 IRGA, the model 42CTL

    chemiluminescence NOx analyzer, and the Shimadzu

    GC-8AIE were automated. More details on these

    analytical measurements and calibration procedures

    are presented elsewhere (Filippa et al. 2009, for the

    GC-8AIE; Helmig et al. 2009, for the 42CTL; Liptzin

    et al. 2009, for the LI-7000 IRGA). The sampling

    manifold, calibration system, and data acquisition

    were controlled through an array of digital input/

    output modules, temperature input components, and

    LabVIEW software (National Instruments, Austin,

    Texas, USA). The GC-ECD signals were acquired

    through a PeakSimple system (SRI Instruments,

    Torrance, California, USA).

    Ancillary data

    Temperatures at each sampling height were measured

    using type-E thermocouples (Omega Engineering,

    Inc., Stamford, Connecticut, USA) that were covered

    by white shrink tubing to reduce radiation artifacts.

    Wind speed, at 6 m above the ground on a

    meteorological (MET) tower 10 m away, was mea-

    sured using a 05103-L R.M. Young Wind Monitor

    (Campbell Scientific, Logan, Utah, USA). Barometric

    pressure was measured using a CS105 Vaisala

    PTB101B Barometer (Campbell Scientific, Logan,

    Utah, USA). Differential pressure between the soil–

    snow and snow–atmosphere interfaces was collected

    at the MET tower using a differential pressure sensor

    (PX277-0.1D5V, Omega Engineering, Inc., Stam-

    ford, Connecticut, USA) with 1 Hz sampling rate and

    1-min averages stored. The high pressure inlet was

    placed at 3 m above the soil surface and the low

    pressure inlet was placed on the ground. Given this

    setup, if pressure at the ground was greater than at the

    snow surface, then the resulting pressure gradient was

    measured as a negative signal.

    Readings of snow depth were conducted at

    1–2 week intervals from calibrated marks on the

    snow tower. Gaps in the snow depth record were

    filled by comparing and interpolating the daily snow

    depth record from the SNOTEL site, 1.5 km from the

    Soddie. An analysis of the linear correlation between

    concurrent snow-depth data from these two sites

    resulted in a r2 = 0.91. These snow depth data were

    Fig. 1 Series ofphotographs showing the

    snowpack sampling tower

    (snow tower). The left

    picture was taken before the

    onset of snowfall in autumn

    2006. The upper-rightpicture shows the snow

    tower buried in *2 m ofsnow with the top sampling

    inlet (at 245 cm) above the

    snowpack on 11 March

    2007. The lower-right photoshows the syringe fiber

    glass filter and the type-E

    thermocouple mounted to

    one of the ends of the

    horizontal cross arms on the

    snow tower

    Biogeochemistry (2009) 95:95–113 99

    123

  • used for flux calculations including the ambient air

    measurement data (245 cm). An effective snowpack

    diffusion distance to the snow surface was deter-

    mined by calculating the distance from the respective,

    fixed height inlet pair to this determined snow surface

    height. Gas measurements from the 245 cm height

    were assumed to be representative for conditions at

    the snow-atmosphere interface. This simplification

    neglects the concentration gradients between the

    snow surface and the 245 cm height.

    Soil moisture was measured using four CS616-L

    Water Content Reflectometers with 30 cm long

    probes installed vertically into the soil in a 1 m

    radius from the snow tower (Campbell Scientific,

    Logan, Utah, USA), following the protocol at the C-1

    Ameriflux site on Niwot Ridge (Bowling et al. 2009),

    Fig. 2 Schematics of winter 2007 season snowpack samplingtower with plumbing and instrument configuration. The sky-blue dotted pattern area is the snowpack. The brown dashedline defines the ground. The instruments are located in anunderground lab. Abbreviations of the cylinder labels in the

    diagram reflect the following: BA = breathing air standard,

    C = 1% CO2 standard, N1 = 320 ppb N2O standard,

    N2 = 408 ppb N2O standard, N3 = 485 ppb N2O standard,

    NO = 100 ppb NO standard, P5 = 5% CH4/95% Ar carrier

    gas (Color figure online)

    100 Biogeochemistry (2009) 95:95–113

    123

  • with the average of the four sensors used for all

    further data applications.

    Snow density was measured in an adjacent open

    meadow area*30 m away from the primary study sitewith similar vegetation and slope characteristics.

    Snow pits were dug every 1–2 weeks, upslope first

    and progressed downhill for each new snow pit. Snow

    properties including depth, density, grain type, grain

    size, and stratigraphy were measured using previously

    published protocols (Williams et al. 1996, 1999).

    Results

    Snow physical properties

    Continuous snowpack at this site typically develops

    between the end of October and early November. The

    seasonal snow cover reached a maximum depth of

    2.2 m in March 2007, and all snow had melted by the

    first week of June (Fig. 3). The snowpack develop-

    ment in 2007 was similar to our five seasons of

    measurements, where seasonal maximum snowpack

    depths of 1.8–2.2 m occurred between day of year

    (DOY) 65 and 85. While there were occasional minor

    reductions in snowpack depth throughout the winter,

    with a more significant one occurring during DOY

    80–85 in 2007, the springtime snow melt with

    significant and continuous snow depth reductions

    did not commence until DOY 130. The physical

    changes associated with this transition are also

    evident in the temperature and density records.

    Snowpack temperatures were usually warmer than

    ambient air temperatures throughout the season

    (Fig. 3a). The average air temperature during the

    snow-covered period was -5.2�C, with minimumvalues of -25�C in mid-winter and an increasingtrend in late winter and early spring. The diurnal and

    daily fluctuations of air temperature progressively

    damped within the first 10–50 cm below the snow

    surface. The temperature at the soil–snow interface

    increased to about 0�C when the snowpack reached adepth of *12 cm and remained as such until the endof the season (Fig. 3a). At DOY 67, the upper

    snowpack layers suddenly warmed. The entire snow-

    pack turned isothermal over an 8 day period. How-

    ever, it took another 55 days from that point until

    significant snow melt occurred.

    Snowpack density generally increased over time

    (Fig. 3b). The initial density was *200 kg m-3 early

    Fig. 3 a A temporal and vertical development of thesnowpack temperature. The black solid line is the snow heightabove the ground (cm). The color bar is the snowpacktemperature (�C). The inverted dark grey triangles denote timewhen actual snow depth measurements were taken. The gaps in

    the snow depth data were filled through extrapolating daily

    snow depth measurements from the nearby SNOTEL site (see

    text for details). b Contour plot of snow density (kg m-3)generated from the biweekly snow pit excavation with 10-cm

    vertical resolution. The inverted green triangles denote thedates when actual snow density samples were taken. Gaps

    between sampling periods were filled through linear interpo-

    lation (Color figure online)

    Biogeochemistry (2009) 95:95–113 101

    123

  • in the season; the maximum density measured of

    500 kg m-3 occurred during the snowmelt period.

    Density normally increased with depth due to com-

    pression caused by the overlying snow. However, at

    depths below 60 cm, density remained lower and

    showed little change over time until active snowmelt.

    This section of the snowpack was composed of kinetic

    grains (depth hoar), which resists compression.

    CO2 and H2O vapor concentrations within the

    snowpack

    A typical plot of the snowpack concentrations of CO2and H2O vapor during a sampling cycle is presented

    in Fig. 4. After 3 min of the transition between

    levels, concentrations of CO2 and H2O vapor

    remained relatively constant for the remainder of

    the measurement interval. Because of the changing

    signal during the transition from one inlet to the next,

    the five points around the transition period were

    omitted for calculating the gas concentration for each

    measurement time interval (Fig. 4).

    CO2 fluxes

    The time series of calculated daily mean fluxes from

    all possible gradient combinations for winter 2007 is

    shown in Fig. 5. These data show a remarkable

    variability both between results from different inlet

    heights and between measurement days. Figure 5 also

    shows the median value calculated from all available

    gradient interval fluxes at a given time. Closer

    inspection of this record revealed that individual

    gradient data typically fall within a range of ±50%

    of the overall median, but it was not uncommon that

    data from a selected gradient deviated up to 100%

    from the median. Results derived from measurements

    at the 0 cm inlet frequently deviated from the other

    inlet data, with gradients between the 0 and 10 cm

    inlets often being unusually small and on occasion

    even negative. This behavior was interpreted as this

    measurement interval being too close to the ground

    and within the surface roughness layer at this site,

    where consistent gradients are not well developed.

    Figure 5 also shows the median value calculated from

    Fig. 4 Example of a timeseries of CO2 and H2O

    vapor concentration data

    during one sampling cycle

    on 11 March 2007 with the

    x-axis scale indicating timefrom the beginning of the

    cycle (2:40 am). The dotsare 1-min averages of data

    collected every 10 s. The

    white dots in each 10-minsampling interval represent

    the data that were averaged

    for determining gas

    concentration gradients

    (qC/qz). The black dotswere omitted from the

    calculation for determining

    the gas concentration

    gradient because the values

    may be compromised

    during the switching

    between sampling inlets.

    The snowpack depth during

    this sampling cycle was

    1.87 m

    102 Biogeochemistry (2009) 95:95–113

    123

  • all available gradient interval fluxes at a given time.

    Our estimate of the diffusive CO2 flux based on the

    seasonal average of the daily mean ranged from 0.58 to

    1.08 lmol m-2 s-1 in 2007, depending on the pairs ofinlets used for the gradient calculation (Fig. 5).

    However, of the 27 gradient combinations, sixteen

    of them gave results between 0.70 and 0.90 lmolm-2 s-1. The linear correlation of the all-season data

    from the 27 gradient combinations illustrated that—

    not unexpectedly—correlations were highest where

    gradient intervals overlap due to the autocorrelation in

    these data. For example, the correlation coefficient for

    gradients calculated from 90 to 60 cm and from 120 to

    30 cm was 0.99. Furthermore, correlations were high

    for data from adjacent gradients (please see Fig. 4 in

    Filippa et al. (2009), which graphically displays the

    corresponding analysis for calculated N2O fluxes from

    the 90–60 cm vs. the 60–30 cm gradient data), and

    became weaker the further the gradients were sepa-

    rated vertically. Also, better agreement was generally

    seen in data from the middle of the snowpack, whereas

    results from sampling near the bottom and involving

    measurements from the ambient inlet often showed

    weaker correlations. The correlation coefficient r was

    only 0.29 between gradients calculated from 30 to

    10 cm and from 120 to 90 cm. In essence, the linear

    correlation results of the fluxes were dictated by the

    concentration gradient, as shown in Eq. (1). Absolute

    correlation values r of 1 or close to 1 between

    different-height gradients implied that the fluxes

    through the snowpack were constant, which is a key

    assumption of the DM. This behavior also implies that

    diffusion was the dominant process in the transport of

    CO2 though the snowpack. Our multiple-inlet gradient

    and flux data illustrate that for the center portion of the

    snowpack, these conditions (i.e., linear concentration

    gradients and dominant diffusion transport) were

    reasonably well met and that the diffusion model

    was an appropriate tool for estimating fluxes. Conse-

    quently, most of our subsequent analyzes were done

    using the data from the 60–30 cm gradient interval.

    These data showed relatively high correlation with

    most non-overlapping inlet pair combination (range in

    r values of 0.29–0.80). Furthermore, the 60–30 cm

    inlet data were the longest and most steady record for

    the 2005–2006 winter season (here on as winter 2006)

    and winter 2007.

    Discussion

    Potential sampling effects

    Active sampling from interstitial air in the snowpack

    represents a disturbance to the natural air transport and

    this can generate artifacts in the measured snowpack

    data (Albert et al. 2002; Domine et al. 2008). We

    evaluated the potential artifact caused by our sampling

    Fig. 5 Daily winter 2007CO2 flux for all sampling

    height combination. The

    thick black line depicts thedaily median flux data

    calculated from the results

    of all gradient combinations

    (Color figure online)

    Biogeochemistry (2009) 95:95–113 103

    123

  • technique using an approach similar to the one

    presented by Bowling et al. (2009). Our sampling

    system pulled *20 L of air from the paired inletsduring each sampling interval. The *10 L of airwithdrawn at each inlet, at a snowpack density of

    400 kg m-3 (which was the seasonal average), and

    assuming a sphere volume, would yield an effective

    radius of *16 cm per inlet of the sampling volume.All, but one, sampling ports were 30 cm apart in the

    vertical, so there is very little overlap within this

    radius. (The closer distance between the 0 and 10 cm

    height inlet pairs was the only exception to this

    requirement). Based on gas-phase diffusive transport

    only (Seinfeld and Pandis 2006, p 551), we estimated

    that it would take *44 min within this volume for airto re-equilibrate to its original condition. This is faster

    than the 80-min cycle used in our experiment. [With

    increased (decreased) snow density, the re-equilibrium

    time would increase (decrease), and the sample will

    reflect a larger (smaller) volume of snowpack]. Data

    from our sampling protocol confirmed the validity of

    these assumptions. As seen in Fig. 4, other than during

    the transition period, there was comparatively little

    change in gas concentrations during the 10-min

    sampling interval. This meant that gas sampled during

    this time remained relatively representative of the

    given inlet height. Also, closer inspection of these

    records revealed that the relative change of signal was

    largest for the 0 cm measurement height, which

    reflects the fact that the available airspace extends

    only in one direction; therefore, air that replaced the

    sampled air volume is expected to be drawn from a

    larger vertical distance. The larger degree of overlap-

    ping sampling volumes for the 0 and 10 cm heights

    explained to some extend the deviating behavior in

    observed gradient from these two heights. In summary,

    our results were in agreement with the findings of

    Bowling et al. (2009) who re-measured several inlets

    25 min after the initial measurement and found no

    difference in CO2 concentrations or isotopes of carbon

    and concluded that the advective influence of their

    sampling under these conditions was negligible.

    Dependence of CO2 flux on environmental

    conditions

    In general, CO2 fluxes gradually increased over the

    snow-covered season, reaching a maximum around the

    onset of snowmelt. An interesting question was what

    drove the daily variations in CO2 (Fig. 5), and if the

    change in flux was driven by changes in the source

    term, i.e., day-to-day changes in the emission of CO2from soil to the snowpack, or to other physical or

    chemical processes that influenced the flux calculation.

    For the investigation of this question, we focused on

    the 30-day time period between DOY -11 and 19 when

    the snow depth was relatively constant at *120 cm(Fig. 6a). Over this time span, the soil volumetric

    water content remained constant at *0.20 m3 m-3

    (Fig. 6b), and the temperature near the soil–snow

    interface was always*0�C (Fig. 6d). In contrast to therelatively constant snow depth, snow temperature, and

    soil moisture content, data for wind speed ranged over

    more than an order of magnitude from\0.2 to 6 m s-1

    (Fig. 6c). In turn, there was a tendency that higher

    wind speeds were associated with lower differential

    pressures (Fig. 6c). Concentrations of CO2 increased

    from ambient levels at the snow surface (* 385 ppm)to *4,000 ppm at the bottom of the snowpack(Fig. 6e). Comparison between the CO2 concentration

    and wind speed time series data suggested that CO2concentrations in the snowpack generally decreased

    with increasing wind speed. For example, between

    DOY -6 and -5, at the 10 cm level, CO2 concentrations

    decreased by almost a factor of 2 from 2,690 to

    1,475 ppm, as wind speed increased from 1.5 to

    5.9 m s-1. As a result, calculated CO2 fluxes

    decreased from 0.67 to 0.16 lmol m-2 s-1 (Fig. 6f).

    Uncertainties in flux estimates using the diffusion

    model

    Snow permeability and gas transport are related to

    snow density (Albert and Hardy 1995; Albert and

    Shultz 2002; Albert et al. 2004) as permeability is

    affected by porosity and tortuosity, even though there

    is no direct relationship between porosity, tortuosity,

    and permeability (Hardy et al. 1995). In our flux

    calculation, gas diffusivity, following Eq. (2), was

    derived indirectly from snow density, as porosity and

    tortuosity were calculated from the density values

    [see Eqs. (3) and (5)]. For example, the diffusion rate

    at a density of 250 kg m-3 was 87% faster than at a

    density of 500 kg m-3, assuming all other variables

    were unchanged. Thus, errors in the estimate of CO2flux through snow caused by incorrect measurements

    of density varied as density changed. Errors in

    measuring the density of snow have been estimated

    104 Biogeochemistry (2009) 95:95–113

    123

  • at 10–15% (Sommerfeld et al. 1996; Hubbard et al.

    2005). In our case, the uncertainty in applied density

    was likely higher, as we applied density data from an

    adjacent plot, and the biweekly density determina-

    tions did not allow for capturing density changes that

    occurred at higher frequency than this sampling

    interval. We evaluated how errors in measuring

    density propagated to errors in estimating the flux

    of CO2 using errors in the measurement of density of

    ±10, 20, and 30%. The calculated error increased

    non-linearly at higher snow densities and with the

    uncertainty in the determination of snow density. For

    instance, a 10% error in the measurement of snow

    density resulted in an error in the estimated CO2 flux

    on the order 5% for a snow density of 300 kg m-3

    but more than 10% at a density of 500 kg m-3. For

    the range of snow densities encountered during the

    experiment at the Soddie sites, we estimated that the

    error in calculating the flux of CO2 ranged from 2 to

    9% (Supplementary Material, Fig. S1). Another

    potential complication in the diffusion flux calcula-

    tion was the presence of crust and ice layers, which

    may constitute diffusion barriers that were not

    considered in the Fick’s law calculation. Snow pit

    analyses showed that ice layers were uncommon at

    the site, as snow surface temperatures at this high

    elevation generally remain below freezing (which

    avoids freeze-thaw events that form crust and ice

    layers) until the onset of the spring melt period. There

    are several reasons why our sampling technique is

    rather robust against the influence of inhomogeneous

    snow density, and ice lenses and crusts. The variation

    Fig. 6 Dataset from DOY-11 to 19. a Snow heightfrom the ground. The greydots are the actual snowdepth measurements. The

    solid line is the interpolateddata based on the daily

    SNOTEL snow depth

    measurements (see text for

    details). b Soil moisturedata. The line is the averageof the four soil moisture

    sensors surrounding the

    snow tower in a 1-m radius.

    c Wind speed (dark bluesolid line) and differentialpressure (sky-blue solidline). d Temperature datafrom the snow tower (in and

    above the snowpack). e CO2concentration in and above

    the snowpack measured

    from the snow tower. The

    green and red arrows markthe start and end when the

    instrument range was

    maxed out at 3000 ppm at

    one or more sampling

    heights. f Calculated fluxusing the 30 and 60 cm inlet

    data without consideration

    of wind correction (Color

    figure online)

    Biogeochemistry (2009) 95:95–113 105

    123

  • of snowpack density was accounted for in the flux

    calculation. For each sampling interval, the snowpack

    density corresponding to that sampling interval and

    snowpack layer was used to calculate the fluxes. This

    resolves the temporal and spatial variations the

    snowpack density would have on the flux. Concep-

    tually, our sampling system in the snowpack was

    similar to a chamber with open sides giving it

    horizontally infinite spatial scale. Since ice lenses and

    crusts do not span over an infinitely large area, gases

    diffusing through the snowpack would continue to

    travel around the ice lenses and crusts, and gas

    concentration at each arbitrary horizontal layer in the

    snowpack would equilibrate and maintain a distinct

    concentration gradient. The deviation in flux results

    from different height intervals affirms (Fig. 5) and

    defines the maximum effect that this error; as already

    elucidated above, these data show that flux results

    generally agreed within a factor of 1.5 from different

    height intervals.

    Errors in the measurement of the snowpack height

    and the height of the tower arms were also a source of

    uncertainty when calculating the flux of CO2. All

    gradient distances within the snowpack were deter-

    mined by the distance between the tower cross arms.

    Due to the pressure exhibited from the compacting

    and creeping snow, some of the cross arms were bent

    by up to 5 cm, which introduces an error of as much

    as 15% in the flux calculation. The flux from any inlet

    to the snow surface depended on the accuracy on the

    snow depth determination. The associated error

    would be variable, depending on the time passed

    from the last visual inspection, and the accuracy of

    the interpolation from the SNOTEL record. Further-

    more, this error relied on the absolute distance of the

    respective inlet to the snow surface. As this error

    propagated, its relative value would be larger for

    inlets closer to the atmosphere than for inlets nearer

    to the soil surface. We estimated the overall range of

    the error contributed from uncertainties in the gradi-

    ent interval to the surface to be on the order of 10 to

    100% in the most extreme cases; the latter for inlets

    very close to the snow surface. Temperature mea-

    surements inside the snowpack are expected to have a

    minor contribution (\5%) to the error estimate, assnowpack air temperatures were measured with a

    relative high accuracy and showed little variation

    with time and space (Fig. 3). This error may increase

    slightly nearer to the surface where temperature

    changes were more pronounced than deeper in the

    snow.

    We also investigated how sampling frequency may

    bias the flux calculation. Among previous studies the

    frequency of measurements varied widely, ranging

    from as often as half-hourly (e.g., Musselman et al.

    2005) to as few as twice per winter (e.g., Hubbard et al.

    2005). We used a bootstrap randomization without

    replacement test to examine the sampling frequency

    effects (Supplementary Material, Fig. S2). First, we

    calculated a daytime (9 a.m.–5 p.m.) mean to provide a

    more realistic comparison to manual sampling proto-

    cols. For each sampling frequency the appropriate

    number of daytime values (ranging from 5 to

    120 days) was randomly selected and a mean calcu-

    lated. This process of calculating means was repeated

    1,000 times, and the 95% confidence interval was

    determined as the 25th and 975th mean (Manly 1997).

    The 95% confidence limit for the continuous data was

    0.69–0.82 lmol m-2 s-1 in 2007. Monthly samplingresulted in an increase in the 95% confidence interval

    ranging from 0.5 to 1.1 lmol m-2 s-1. The 95%confidence interval for weekly sampling (30 samples)

    of 0.62–0.88 lmol m-2 s-1 moved close to the resultsfrom our continuous measurements. The results from

    this bootstrap analysis meant that a weekly sampling

    frequency (i.e. one gradient per week) would give a

    mean flux value that was within ±10% of the results

    from our continuous measurements of CO2 flux.

    Assessment of advection on flux calculations

    The observed day-to-day variability in calculated

    fluxes and differences in results from different

    gradient intervals fully cannot be accounted for even

    when combing all of the above considered sampling

    and measurement errors and uncertainties, conse-

    quently other factors must contribute to the flux

    variability. Other research has shown that the CO2efflux from snow-covered soils is sensitive to volu-

    metric soil moisture and soil temperature (e.g. Oechel

    et al. 1997; Winston et al. 1997; Groffman et al.

    2006; Monson et al. 2006). The lack of significant

    changes in these variables during DOY -11 to 19

    (Fig. 6) suggests that CO2 emissions from the soil

    might be reasonably constant over this 30 day period;

    it is therefore also unlikely that the variability in the

    flux data is caused by changes in the CO2 source

    term. We, therefore, conducted further analysis of the

    106 Biogeochemistry (2009) 95:95–113

    123

  • potential role that advection may play in the

    measured CO2 concentrations and calculated diffu-

    sion fluxes in the snowpack.

    Snowpack ventilation has been shown to be related

    directly to the high frequency pressure gradients that

    develop above the surface and inside the snowpack

    during windy conditions (Albert and Hawley 2002;

    Massman 2006; Massman and Frank 2006). From the

    differential pressure data, measured at 1-s resolution,

    the 10-min mean values (Fig. 7a), the standard

    deviation within each 10-min window (Fig. 7b), and

    the amplitude (Fig. 7c) were calculated for the

    corresponding time window. All three of these graphs

    show strong correlations between these variables,

    with the highest degree of correlation seen in the

    dependency of the standard deviation of the differ-

    ential pressure data on the 10-min mean wind speed

    (Fig. 7b). Waddington et al. (1996) described that the

    forcing term, the pressure amplitude, has a 2nd order

    relationship with wind speed (Aop/ �u2, where A

    opis

    the amplitude of the mean differential pressure and �u

    is the mean wind speed). However, depending on the

    topography of the surrounding terrain, a 1st order

    relationship between pressure amplitude and wind

    speed Aop/ �u

    � �may be more appropriate (Massman

    and Frank 2006). Our regression analyses, examining

    different regression fits, resulted in a best fit (highest

    r2) for the 1st order relationship between differential

    pressure amplitude and 10-min mean wind speed

    (Fig. 7c). These analyses showed that the pressure

    amplitude was directly proportional to wind speed.

    Thus, it appeared appropriate to use wind speed as the

    forcing proxy for assessing advection.

    In Figure 8a, concentration gradients from the

    60–30 cm interval were plotted against the mean

    hourly wind speed during the measurement interval.

    With increasing wind speed, the magnitude of the CO2concentration gradient oCCO2=ozj j became progres-sively smaller. The concentration gradient reached

    -141 ppm m-1 at the highest recorded winds, while

    the gradient for zero wind conditions, as extrapolated

    from the 2nd order polynomial regression fit, was

    -1,090 ppm m-1. Concentration gradients varied by

    a factor of *8 under the wind conditions encounteredduring this experiment. As the concentration gradient

    translated proportionally to the final CO2 flux, the flux

    results shown in Fig. 8b display a similar range of

    dependency, with calculated flux values at the highest

    Fig. 7 Relationship between measures of the pressure fluctu-ations, representing the effect of wind pumping, as a function

    of absolute 10-min average wind. The differential pressure data

    represent a the mean of 1-s data, b the standard deviation of the4,464 measurements within each 10-min interval, and c themean value of the amplitudes that were calculated from the

    recorded maximum and minimum values during each minute.

    Regression curves shown are the selected best fit results of 1st

    and 2nd order regression calculations that were examined. The

    blue curves represent the 95% confidence interval of theregression result (Color figure online)

    Biogeochemistry (2009) 95:95–113 107

    123

  • winds dropping to a mere 1/8th of values derived under

    calm conditions.

    Some of our earlier analyses of the diurnal depen-

    dency of the CO2 flux had suggested on average 25%

    flux increase between early morning and mid afternoon

    hours. Figure 9 illustrates this analysis by showing the

    mean hourly deviation from the daily mean flux values

    (for winters 2006 and 2007) that led to this assumption.

    The added wind speed data shed new light on this

    discussion. There was an obvious anti-correlation in

    the diurnal behavior of winds and calculated flux

    values, with higher winds coinciding with lower flux

    values and vice versa. Using the amplitudes of winter

    2007 data series, the 0.3 m s-1 change in wind speed

    corresponded to a -0.05 lmol m-2 s-1 change in thediffusion-calculated CO2 flux. Similarly, for winter

    2006, a 0.3 m s-1 change in wind speed yielded a

    0.03 lmol m-2 s-1 change in CO2 flux. The winter2007 result agreed with Fig. 10b, where similarly the

    slope of the regression function within the linear range

    predicted a -0.06 lmol m-2 s-1 change in the diffu-sion-CO2 flux for a 0.3 m s

    -1 wind speed increase.

    Since results presented in Figs. 8b and 9 were in

    qualitative agreement, results from the 30 day period

    of analysis shown in Fig. 8b are expected to be

    representative for the entire season shown in Fig. 9.

    Bowling et al. (2009) illustrated that the transport

    of trace gases through the snowpack at the C-1

    Fig. 8 a CO2 concentration gradient (oCCO2=oz) and b CO2flux derived from inlets at 30 and 60 cm and wind speed from

    DOY -11 to 19 with 2nd order polynomial fit solutions

    (y = cx2 ? bx ? a). The blue curves depict the 95%

    confidence interval for the fits. The y-intercept, a term, wouldrepresent the CO2 concentration gradient and CO2 flux under

    zero wind condition (Color figure online)

    Fig. 9 Deviation fromdaily means of wind speed

    (left y-axis) and CO2 flux(right y-axis) for winter2006 and 2007. The errorbars are the standard errorsfor each bin of data. The

    CO2 fluxes were calculated

    from the 60–30 cm

    concentration gradient for

    both winter 2006 and 2007

    108 Biogeochemistry (2009) 95:95–113

    123

  • Ameriflux site on Niwot Ridge was dominated by

    diffusion with short-term variability driven by advec-

    tion. Furthermore, they showed that the influence of

    diffusion was differentially apparent at greater depths.

    We evaluated how advection may change the concen-

    trations of CO2 between the soil surface and the top of

    the snowpack by testing the relationship between CO2concentrations at each inlet height and wind speed

    with 2nd order polynomial solutions (y = cx2 ? bx

    ? a) [Fig. 10]. The regression term a, denoting the

    zero-wind speed snowpack CO2 concentration at each

    inlet height, linearly increased going from the snow

    surface towards the bottom of the snowpack. The term

    b demonstrates the ‘‘strength’’ of the wind’s influence

    on the CO2 concentration. This value steadily

    increased towards the soil surface (going deeper into

    the snow), indicating that the wind effect on absolute

    CO2 levels was highest at the bottom of the snowpack,

    where absolute CO2 concentrations were the highest.

    Solutions for c characterized the curvature of the best

    fit equation. Here, values increased towards the bottom

    of the snowpack, indicating an increasingly higher

    sensitivity towards winds deeper in the snow. r2

    values, denoting the quality of the regression fit,

    increased with increasing depth, showing again that

    the wind pumping effect was best described by these

    algorithms at the bottom of the snowpack. Lastly, the

    95% confidence interval demonstrated how at higher

    wind speeds, the uncertainty of CO2 concentration in

    the snowpack increases. An important conclusion

    from these analyses is that measurements made during

    high wind conditions would have increased errors in

    calculating CO2 gradients and for flux calculations.

    Similar to Bowling et al. (2009), our findings showed

    that advection effects at the bottom of the snowpack

    were described best by these algorithms.

    To demonstrate the bias from wind pumping on

    flux results, CO2 fluxes calculated from concentration

    data, as obtained during actual wind speed conditions

    encountered during DOY -11 to 19, were compared

    with fluxes derived from extrapolated gradients for

    zero wind speed conditions. These latter values were

    derived from the regression equations (intercept

    values of the regression solutions presented in

    Fig. 10). Fluxes were calculated in both ways for

    every sampling interval combination. While results

    for individual gradients showed a fair degree of

    variability, overall the findings shown in Fig. 11

    illustrate that in every case higher flux values were

    obtained for the wind-corrected calculation. The

    median flux value calculated from the in situ gradient

    data of 0.54 lmol m-2 s-1 accounts to 64% of therespective value derived from the wind-corrected

    calculation (0.85 lmol m-2 s-1). This findingdefines the possible error that is associated with the

    application of Fick’s law under neglect of wind

    pumping effects and the resulting underestimation of

    flux results.

    Summary and conclusions

    We have evaluated the DM method for studying

    wintertime CO2 fluxes at Niwot Ridge, Colorado. A

    particular advantage of this method stems from the

    reduced transfer rate (in comparison to ambient air) in

    the snowpack, resulting in largely elevated concentra-

    tions that build up inside the snow. This condition

    causes concentration gradients of gases released from

    the subniveal soil reach magnitudes that are much

    larger than above the surface, which allowed for their

    measurement with relatively simple analytical instru-

    ments. Deriving fluxes of these gases requires quan-

    titative description of the gas transport in the airspace

    of the snowpack. Here, we used the by far most

    extensive dataset published to date to examine the

    commonly applied Fick’s law of diffusion approach

    for computation of fluxes from these data.

    Fluxes calculated from all possible gradients from

    eight inlet heights were found to generally agree

    within a factor of 1.5 with the overall median. Several

    factors contribute to variability and uncertainty in

    these data, with the error in the snow porosity and

    tortuosity variables, which both are indirectly derived

    from snow density measurements, being one of the

    major contributing factors. This research demon-

    strates how these uncertainties are reduced when data

    from several gradient heights can be statistically

    evaluated and combined for deriving a more repre-

    sentative, average flux value.

    Fig. 10 Relationship between CO2 concentration measured ateach inlet height above and in the snowpack and wind speed

    from DOY -11 to 19 with 2nd order polynomial fit

    (y = cx2 ? bx ? a). The blue curve shows the 95% confidenceinterval for the 2nd order fit. Coefficient values printed in red

    denote cases that are statistically significant (p \ 0.05). They-intercept, a term, represents the extrapolated CO2 concen-tration under zero wind condition (Color figure online)

    c

    Biogeochemistry (2009) 95:95–113 109

    123

  • 110 Biogeochemistry (2009) 95:95–113

    123

  • Using wind and differential pressure measurements,

    we were able to develop the constraints from the

    neglect of wind pumping on fluxes derived by the

    diffusion approach. Gas concentrations of soil emis-

    sions and their snowpack gradients in general declined

    with increasing winds, causing a negative bias on

    calculated fluxes, with fluxes calculated at the highest

    observed winds being less than 13% of those derived

    under no-wind conditions. Using an actual 3-week

    record from mid-winter, and averaging over all possi-

    ble gradient combinations, it was determined that for

    conditions at this particular site fluxes calculated by the

    diffusion method, neglecting wind-pumping effects,

    were 36% lower than wind-corrected calculations.

    Acknowledgments This research at Niwot Ridge, Coloradowas funded by the Long-Term Ecological Research grant from

    the National Science Foundation (award# NSF DEB9211776).

    This work was also supported by NSF grant OPP-0240976. B.

    Seok acknowledges the Biosphere-Atmosphere Research

    Training (BART) fellowship from the NSF-IGERT program

    administrated by the University of Michigan. We thank W.

    Massman, US Forest Service, Fort Collins, for fruitful

    discussions throughout this study, M. Losleben, University of

    Colorado at Boulder, for help with site maintenance and

    logistics, and many other University of Colorado colleagues for

    support and encouragement in the Niwot Ridge research

    projects. Any opinions, findings, and conclusions expressed in

    this material are those of the authors and do not necessarily

    reflect the views of the National Science Foundation.

    Open Access This article is distributed under the terms of theCreative Commons Attribution Noncommercial License which

    permits any noncommercial use, distribution, and reproduction

    in any medium, provided the original author(s) and source are

    credited.

    References

    Albert MR, Hardy JP (1995) Ventilation experiments in a

    seasonal snow cover. In: Tonnessen KA, Williams MW,

    Tranter M (eds) Biogeochemistry of seasonally snow-

    covered catchments. IAHS Press, Wallingford, pp 41–49

    Albert MR, Hawley RL (2002) Seasonal changes in snow

    surface roughness characteristics at Summit, Greenland:

    implications for snow and firn ventilation. Ann Glaciol

    35:510–514. doi:10.3189/172756402781816591

    Albert MR, Shultz EF (2002) Snow and firn properties and air-

    snow transport processes at Summit, Greenland. Atmos

    Environ 36:2789–2797. doi:10.1016/S1352-2310(02)001

    19-X

    Albert MR, Grannas AM, Bottenheim J, Shepson PB, Perron

    FE (2002) Processes and properties of snow-air transfer

    in the high Arctic with application to interstitial ozone

    at Alert, Canada. Atmos Environ 36:2779–2787. doi:

    10.1016/S1352-2310(02)00118-8

    Albert M, Shuman C, Courville Z, Bauer R, Fahnestock M,

    Scambos T (2004) Extreme firn metamorphism: impact of

    decades of vapor transport on near-surface firn at a low-

    accumulation glazed site on the East Antarctic plateau. Ann

    Glaciol 39:73–78. doi:10.3189/172756404781814041

    Bowling DR, Massman WJ, Schaeffer SM, Burns SP, Monson

    RK, Williams MW (2009) Biological and physical influ-

    ences on the carbon isotope content of CO2 in a subalpine

    forest snowpack, Niwot Ridge, Colorado

    Brooks PD, Williams MW (1999) Snowpack controls on nitro-

    gen cycling and export in seasonally snow-covered catch-

    ments. Hydrol Process 13:2177–2190. doi:10.1002/(SICI)

    1099-1085(199910)13:14/15\2177::AID-HYP850[3.0.CO;2-V

    Fig. 11 Calculated fluxunder observed and inferred

    zero wind conditions for

    different concentration

    gradient intervals (category

    axis) from DOY -11 to 19.

    The wind-corrected

    condition is derived from

    the [CO2] versus wind

    speed relationship shown in

    Fig. 10. The medians for

    each scenario (observed and

    wind-corrected) are shown

    in its respective coloredsolid lines. For observedwind, the median flux is

    0.54 lmol m-2 s-1. Themedian wind-corrected flux

    is 0.85 lmol m-2 s-1

    (Color figure online)

    Biogeochemistry (2009) 95:95–113 111

    123

    http://dx.doi.org/10.3189/172756402781816591http://dx.doi.org/10.1016/S1352-2310(02)00119-Xhttp://dx.doi.org/10.1016/S1352-2310(02)00119-Xhttp://dx.doi.org/10.1016/S1352-2310(02)00118-8http://dx.doi.org/10.3189/172756404781814041http://dx.doi.org/10.1002/(SICI)1099-1085(199910)13:14/15%3c2177::AID-HYP850%3e3.0.CO;2-Vhttp://dx.doi.org/10.1002/(SICI)1099-1085(199910)13:14/15%3c2177::AID-HYP850%3e3.0.CO;2-Vhttp://dx.doi.org/10.1002/(SICI)1099-1085(199910)13:14/15%3c2177::AID-HYP850%3e3.0.CO;2-V

  • Brooks PD, Williams MW, Schmidt SK (1996) Microbial

    activity under alpine snowpacks, Niwot Ridge, Colorado.

    Biogeochemistry 32:93–113. doi:10.1007/BF00000354

    Domine F, Albert M, Huthwelker T, Jacobi HW, Kokhanovsky

    AA, Lehning M, Picard G, Simpson WR (2008) Snow

    physics as relevant to snow photochemistry. Atmos Chem

    Phys 8:171–208

    Duplessis JP, Masliyah JH (1991) Flow through isotropic

    granular porous-media. Transp Porous Media 6:207–221

    Eisenberg D, Kauzmann W (1969) The structure and properties

    of water. Oxford University Press, Oxford

    Erickson TA (2004) Development and application of geosta-

    tistical methods to modeling spatial variation in snowpack

    properties, Front Range, Colorado. University of Colorado

    at Boulder, Boulder, p 190

    Fahnestock JT, Jones MH, Brooks PD, Walker DA, Welker JM

    (1998) Winter and early spring CO2 efflux from tundra

    communities of northern Alaska. J Geophys Res D

    103:29023–29027

    Filippa G, Freppaz M, Williams MW, Helmig D, Liptzin D,

    Seok B, Hall B, Chowanski K (2009) Winter and summer

    nitrous oxide and nitrogen oxides fluxes from a seasonally

    snow-covered subalpine meadow at Niwot Ridge, Colo-

    rado. Biogeochemistry. doi:10.1007/s10533-009-9304-1

    Groffman PM, Hardy JP, Driscoll CT, Fahey TJ (2006) Snow

    depth, soil freezing, and fluxes of carbon dioxide, nitrous

    oxide and methane in a northern hardwood forest. Glob

    Chang Biol 12:1748–1760. doi:10.1111/j.1365-2486.2006.

    01194.x

    Hardy JP, Davis RE, Winston GC (1995) Evolution of factors

    affecting gas transmissivity of snow in the boreal forest.

    In: Tonnessen KA, Williams MW, Tranter M (eds) Bio-

    geochemistry of seasonally snow-covered catchments.

    IAHS Press, Wallingford, pp 51–59

    Helmig D, Seok B, Williams MW, Hueber J, Sanford JRL

    (2009) Fluxes and chemistry of nitrogen oxides in the

    Niwot Ridge, Colorado snowpack. Biogeochemistry. doi:

    10.1007/s10533-009-9312-1

    Hubbard RM, Ryan MG, Elder K, Rhoades CC (2005) Sea-

    sonal patterns in soil surface CO2 flux under snow cover

    in 50 and 300 year old subalpine forests. Biogeochemistry

    73:93–107. doi:10.1007/s10533-004-1990-0

    Jones HG, Pomeroy JW, Davies TD, Tranter M, Marsh P (1999)

    CO2 in Arctic snow cover: landscape form, in-pack gas

    concentration gradients, and the implications for the esti-

    mation of gaseous fluxes. Hydrol Process 13:2977–2989.

    doi:10.1002/(SICI)1099-1085(19991230)13:18\2977::AID-HYP12[3.0.CO;2-#

    Liptzin D, Helmig D, Williams MW, Seok B, Filippa G,

    Chowanski K, Huber J (2009) Process level control on

    CO2 fluxes from a seasonally snow-covered subalpine

    meadow soil, Niwot Ridge, Colorado. Biogeochemistry.

    doi:10.1007/s10533-009-9303-2

    Manly BFJ (1997) Randomization, bootstrap and Monte Carlo

    methods in biology. Chapman & Hall, London

    Mariko S, Nishimura N, Mo WH, Matsui Y, Kibe T, Koizumi

    H (2000) Winter CO2 flux from soil and snow surfaces in

    a cool-temperate deciduous forest, Japan. Ecol Res

    15:363–372. doi:10.1046/j.1440-1703.2000.00357.x

    Massman WJ (1998) A review of the molecular diffusivities of

    H2O, CO2, CH4, CO, O-3, SO2, NH3, N2O, NO, and

    NO2 in air, O-2 and N-2 near STP. Atmos Environ

    32:1111–1127. doi:10.1016/S1352-2310(97)00391-9

    Massman WJ (2006) Advective transport of CO2 in permeable

    media induced by atmospheric pressure fluctuations: 1.

    An analytical model. J Geophys Res-Biogeosci 111:14

    Massman WJ, Frank JM (2006) Advective transport of CO2 in

    permeable media induced by atmospheric pressure fluc-

    tuations: 2. Observational evidence under snowpacks.

    J Geophys Res-Biogeosci 111:11

    Massman WJ, Sommerfeld RA, Mosier AR, Zeller KF, Hehn

    TJ, Rochelle SG (1997) A model investigation of turbu-

    lence-driven pressure-pumping effects on the rate of dif-

    fusion of CO2, N2O, and CH4 through layered snowpacks.

    J Geophys Res D 102:18851–18863

    Mast MA, Wickland KP, Striegl RT, Clow DW (1998) Winter

    fluxes of CO2 and CH4 from subalpine soils in Rocky

    Mountain National Park, Colorado. Global Biogeochem

    Cycles 12:607–620. doi:10.1029/98GB02313

    McDowell NG, Marshall JD, Hooker TD, Musselman R (2000)

    Estimating CO2 flux from snowpacks at three sites in the

    Rocky Mountains. Tree Physiol 20:745–753

    Monson RK, Sparks JP, Rosenstiel TN, Scott-Denton LE,

    Huxman TE, Harley PC, Turnipseed AA, Burns SP,

    Backlund B, Hu J (2005) Climatic influences on net

    ecosystem CO2 exchange during the transition from

    wintertime carbon source to springtime carbon sink in a

    high-elevation, subalpine forest. Oecologia 146:130–147.

    doi:10.1007/s00442-005-0169-2

    Monson RK, Burns SP, Williams MW, Delany AC, Weintraub

    M, Lipson DA (2006) The contribution of beneath-snow

    soil respiration to total ecosystem respiration in a high-

    elevation, subalpine forest. Global Biogeochem Cycles

    20:13. doi:10.1029/2005GB002684

    Musselman RC, Massman WJ, Frank JM, Korfmacher JL

    (2005) The temporal dynamics of carbon dioxide under

    snow in a high elevation rocky mountain subalpine forest

    and meadow. Arct Antarct Alp Res 37:527–538. doi:

    10.1657/1523-0430(2005)037[0527:TTDOCD]2.0.CO;2

    Oechel WC, Vourlitis G, Hastings SJ (1997) Cold season CO2

    emission from arctic soils. Global Biogeochem Cycles

    11:163–172. doi:10.1029/96GB03035

    Roehm CL, Roulet NT (2003) Seasonal contribution of CO2

    fluxes in the annual C budget of a northern bog. Global

    Biogeochem Cycles 17:9. doi:10.1029/2002GB001889

    Seastedt TR (2001) Soils. In: Bowman WD, Seastedt TR (eds)

    Structure and function of an Alpine ecosystem: Niwot

    Ridge, Colorado. Oxford University Press, New York, pp

    157–173

    Seinfeld JH, Pandis SN (2006) Atmospheric chemistry and

    physics, 2nd edn. Wiley, Hoboken

    Sommerfeld RA, Mosier AR, Musselman RC (1993) CO2,

    CH4 and N2O flux through a wyoming snowpack and

    implications for global budgets. Nature 361:140–142. doi:

    10.1038/361140a0

    Sommerfeld RA, Massman WJ, Musselman RC, Mosier AR

    (1996) Diffusional flux of CO2 through snow: spatial and

    temporal variability among alpine-subalpine sites. Global

    Biogeochem Cycles 10:473–482. doi:10.1029/96GB01610

    Soil Survey Staff (2006) Keys to soil taxonomy, 10th edn.

    USDA-Natural Resources Conservation Service,

    Washington

    112 Biogeochemistry (2009) 95:95–113

    123

    http://dx.doi.org/10.1007/BF00000354http://dx.doi.org/10.1007/s10533-009-9304-1http://dx.doi.org/10.1111/j.1365-2486.2006.01194.xhttp://dx.doi.org/10.1111/j.1365-2486.2006.01194.xhttp://dx.doi.org/10.1007/s10533-009-9312-1http://dx.doi.org/10.1007/s10533-004-1990-0http://dx.doi.org/10.1002/(SICI)1099-1085(19991230)13:18%3c2977::AID-HYP12%3e3.0.CO;2-#http://dx.doi.org/10.1002/(SICI)1099-1085(19991230)13:18%3c2977::AID-HYP12%3e3.0.CO;2-#http://dx.doi.org/10.1007/s10533-009-9303-2http://dx.doi.org/10.1046/j.1440-1703.2000.00357.xhttp://dx.doi.org/10.1016/S1352-2310(97)00391-9http://dx.doi.org/10.1029/98GB02313http://dx.doi.org/10.1007/s00442-005-0169-2http://dx.doi.org/10.1029/2005GB002684http://dx.doi.org/10.1657/1523-0430(2005)037[0527:TTDOCD]2.0.CO;2http://dx.doi.org/10.1029/96GB03035http://dx.doi.org/10.1029/2002GB001889http://dx.doi.org/10.1038/361140a0http://dx.doi.org/10.1029/96GB01610

  • Suzuki S, Ishizuka S, Kitamura K, Yamanoi K, Nakai Y (2006)

    Continuous estimation of winter carbon dioxide efflux

    from the snow surface in a deciduous broadleaf forest.

    J Geophys Res D 111:9

    Takagi K, Nomura M, Ashiya D, Takahashi H, Sasa K,

    Fujinuma Y, Shibata H, Akibayashi Y, Koike T (2005)

    Dynamic carbon dioxide exchange through snowpack by

    wind-driven mass transfer in a conifer-broadleaf mixed

    forest in northernmost Japan. Global Biogeochem Cycles

    19:10. doi:10.1029/2004GB002272

    VanBochove E, Jones HG, Pelletier F, Prevost D (1996)

    Emission of N2O from agricultural soil under snow cover:

    a significant part of N budget. Hydrol Process 10:1545–

    1549. doi:10.1002/(SICI)1099-1085(199611)10:11\1545::AID-HYP492[3.0.CO;2-0

    Waddington ED, Cunningham J, Harder S (1996) The effects

    of snow ventilation on chemical concentrations. In: Wolff

    EW, Bales RC (eds) Chemical exchange between the

    atmosphere and polar snow. Springer, Berlin, pp 403–452

    Welker JM, Fahnestock JT, Jones MH (2000) Annual CO2 flux

    in dry and moist arctic tundra: field responses to increases

    in summer temperatures and winter snow depth. Clim

    Change 44:139–150. doi:10.1023/A:1005555012742

    Williams MW, Helmig D, Blanken P (2009) Introduction:

    white on green: under-snow microbial processes and trace

    gas fluxes through snow, Niwot Ridge, Colorado Front

    Range. Biogeochemistry (this issue)

    Williams MW, Brooks PD, Mosier A, Tonnessen KA (1996)

    Mineral nitrogen transformations in and under seasonal

    snow in a high-elevation catchment in the Rocky Moun-

    tains, United States. Water Resour Res 32:3161–3171.

    doi:10.1029/96WR02240

    Williams MW, Brooks PD, Seastedt T (1998) Nitrogen and

    carbon soil dynamics in response to climate change in a

    high-elevation ecosystem in the Rocky Mountains, USA.

    Arct Alp Res 30:26–30. doi:10.2307/1551742

    Williams MW, Sommerfeld R, Massman S, Rikkers M (1999)

    Correlation lengths of meltwater flow through ripe

    snowpacks, Colorado Front Range, USA. Hydrol Process

    13:1807–1826. doi:10.1002/(SICI)1099-1085(199909)13:

    12/13\1807::AID-HYP891[3.0.CO;2-UWilliams MW, Seibold C, Chowanski K (2009) Storage and

    release of solutes from a seasonal snowpack : soil and

    stream water response, Niwot Ridge, Colorado. Biogeo-

    chemistry. doi:10.1007/s10533-009-9288-x

    Winston GC, Sundquist ET, Stephens BB, Trumbore SE (1997)

    Winter CO2 fluxes in a boreal forest. J Geophys Res D

    102:28795–28804

    Yashiro Y, Mariko S, Koizumi H (2006) Emission of nitrous

    oxide through a snowpack in ten types of temperate

    ecosystems in Japan. Ecol Res 21:776–781. doi:

    10.1007/s11284-006-0174-x

    Zimov SA, Davidov SP, Voropaev YV, Prosiannikov SF,

    Semiletov IP, Chapin MC, Chapin FS (1996) Siberian

    CO2 efflux in winter as a CO2 source and cause of sea-

    sonality in atmospheric CO2. Clim Change 33:111–120.

    doi:10.1007/BF00140516

    Biogeochemistry (2009) 95:95–113 113

    123

    http://dx.doi.org/10.1029/2004GB002272http://dx.doi.org/10.1002/(SICI)1099-1085(199611)10:11%3c1545::AID-HYP492%3e3.0.CO;2-0http://dx.doi.org/10.1002/(SICI)1099-1085(199611)10:11%3c1545::AID-HYP492%3e3.0.CO;2-0http://dx.doi.org/10.1023/A:1005555012742http://dx.doi.org/10.1029/96WR02240http://dx.doi.org/10.2307/1551742http://dx.doi.org/10.1002/(SICI)1099-1085(199909)13:12/13%3c1807::AID-HYP891%3e3.0.CO;2-Uhttp://dx.doi.org/10.1002/(SICI)1099-1085(199909)13:12/13%3c1807::AID-HYP891%3e3.0.CO;2-Uhttp://dx.doi.org/10.1007/s10533-009-9288-xhttp://dx.doi.org/10.1007/s11284-006-0174-xhttp://dx.doi.org/10.1007/BF00140516

    An automated system for continuous measurements of trace gas fluxes through snow: an evaluation of the gas diffusion method at a subalpine forest site, Niwot Ridge, ColoradoAbstractIntroductionDiffusion model (DM) measurement technique

    Materials and methodsStudy siteTrace gas sampling systemAncillary data

    ResultsSnow physical propertiesCO2 and H2O vapor concentrations within the snowpackCO2 fluxes

    DiscussionPotential sampling effectsDependence of CO2 flux on environmental conditionsUncertainties in flux estimates using the diffusion modelAssessment of advection on flux calculations

    Summary and conclusionsAcknowledgmentsReferences

    /ColorImageDict > /JPEG2000ColorACSImageDict > /JPEG2000ColorImageDict > /AntiAliasGrayImages false /DownsampleGrayImages true /GrayImageDownsampleType /Bicubic /GrayImageResolution 150 /GrayImageDepth -1 /GrayImageDownsampleThreshold 1.50000 /EncodeGrayImages true /GrayImageFilter /DCTEncode /AutoFilterGrayImages true /GrayImageAutoFilterStrategy /JPEG /GrayACSImageDict > /GrayImageDict > /JPEG2000GrayACSImageDict > /JPEG2000GrayImageDict > /AntiAliasMonoImages false /DownsampleMonoImages true /MonoImageDownsampleType /Bicubic /MonoImageResolution 600 /MonoImageDepth -1 /MonoImageDownsampleThreshold 1.50000 /EncodeMonoImages true /MonoImageFilter /CCITTFaxEncode /MonoImageDict > /AllowPSXObjects false /PDFX1aCheck false /PDFX3Check false /PDFXCompliantPDFOnly false /PDFXNoTrimBoxError true /PDFXTrimBoxToMediaBoxOffset [ 0.00000 0.00000 0.00000 0.00000 ] /PDFXSetBleedBoxToMediaBox true /PDFXBleedBoxToTrimBoxOffset [ 0.00000 0.00000 0.00000 0.00000 ] /PDFXOutputIntentProfile (None) /PDFXOutputCondition () /PDFXRegistryName (http://www.color.org?) /PDFXTrapped /False

    /Description >>> setdistillerparams> setpagedevice